This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The category of reduced imaginary Verma modules

Juan Camilo Arias, Vyacheslav Futorny and André de Oliveira Institute of Mathematics and Statistics, University of São Paulo, São Paulo, BRAZIL. [email protected] Shenzhen International Center for Mathematics, Southern University of Science and Technology, China and Institute of Mathematics and Statistics, University of São Paulo, São Paulo, BRAZIL. [email protected] Institute of Mathematics and Statistics, University of São Paulo, São Paulo, BRAZIL. [email protected]
Abstract.

For an arbitrary affine Lie algebra we study an analog of the category 𝒪\mathcal{O} for the natural Borel subalgebra and zero central charge. We show that such category is semisimple having the reduced imaginary Verma modules as its simple objects. This generalizes the result of Cox, Futorny, Misra in the case of affine 𝔰𝔩2\mathfrak{sl}_{2}.

Key words and phrases:
2020 Mathematics Subject Classification:
Primary 17B10, 17B67, 17B22

1. Introduction

Let A=(aij)0i,jNA=(a_{ij})_{0\leq i,j\leq N} be a generalized affine Cartan matrix over \mathbb{C} with associated affine Lie algebra 𝔤^\hat{\mathfrak{g}} and Cartan subalgebra 𝔥^\hat{\mathfrak{h}}. Let Π={α0,α1,,αN}\Pi=\{\alpha_{0},\alpha_{1},\cdots,\alpha_{N}\} be the set of simple roots, δ\delta the indivisible imaginary root and Δ\Delta the root system of 𝔤^\hat{\mathfrak{g}}. A subset SΔS\subseteq\Delta is a closed partition if for any α,βS\alpha,\beta\in S and α+βΔ\alpha+\beta\in\Delta then α+βS\alpha+\beta\in S, Δ=S(S)\Delta=S\cup(-S) and S(S)=S\cap(-S)=\emptyset. The classification of closed partitions for root system of affine Lie algebras was obtained by H. Jakobsen and V. Kac in [9] and [10] and independently by V. Futorny in [3] and [5]. They show that closed partitions are parameterized by subsets XΠX\subseteq\Pi and that (contrary to what happens in the finite case) there exists a finite number (greater than 1) of inequivalent Weyl group orbits of closed partitions. When X=ΠX=\Pi we get that S=Δ+S=\Delta_{+} and we can developed the standard theory of Verma modules, but in the case XΠX\subsetneq\Pi we obtain new Verma-type modules called non-standard Verma modules.

The theory of non-standard Verma modules was initiated by V. Futorny in [4] (see also [6]) in the case X=X=\varnothing and continued by B. Cox in [1] for arbitrary XΠX\subsetneq\Pi. The case X=X=\varnothing give rise to the natural Borel subalgebra associated to the natural partition Δnat={α+nδ|αΔ0,+,n}{kδ|k>0}\Delta_{\textrm{nat}}=\{\alpha+n\delta\ |\ \alpha\in\Delta_{0,+}\ ,\ n\in\mathbb{Z}\}\cup\{k\delta\ |\ k\in\mathbb{Z}_{>0}\}. The Verma module M(λ)M(\lambda), of highest weight λ\lambda, induced by the natural Borel subalgebra is called imaginary Verma module for 𝔤^\hat{\mathfrak{g}}, when it is not irreducible it has an irreducible quotient called reduced imaginary Verma module. Unlike the standard Verma modules, imaginary Verma modules contain both finite and infinite dimensional weight spaces. Similar results hold for more general non-standard Verma modules.

In [2], while studying crystal bases for reduced imaginary Verma modules of 𝔰𝔩2^\hat{\mathfrak{sl}_{2}}, it was consider a suitable category of modules, denoted 𝒪red,im\mathcal{O}_{red,im}, with the properties that any module in this category is a reduced imaginary Verma module or it is a direct sum of these modules. In this paper, by appropriate modifications we first define a category 𝒪red,im\mathcal{O}_{red,im} for any affine Lie algebra and we show that all irreducible modules in this category are reduced imaginary Verma modules and, moreover, that any arbitrary module in 𝒪red,im\mathcal{O}_{red,im} is a direct sum of reduced imaginary Verma modules.

It should be noted that the results presented in this paper hold for both untwisted and twisted affine Lie algebras. The paper is organized as follows. In Sections 22 and 33, we define, set the notations and summarize the basic results for affine algebras, closed partitions and imaginary Verma modules. In section 44 we introduce the category 𝒪red,im\mathcal{O}_{red,im} and present some of its properties. Finally, in section 55 we present the main results of this paper.

2. Preliminaries

In this section we fixed some notation and the preliminaries about affine algebras and root datum are set up.

2.1. Affine algebras

Let A=(aij)0i,jNA=(a_{ij})_{0\leq i,j\leq N} be a generalized affine Cartan matrix over \mathbb{C} with associated affine Lie algebra 𝔤^\hat{\mathfrak{g}}. Let D=diag(d0,,dN)D=diag(d_{0},\ldots,d_{N}) be a diagonal matrix with relatively primes integer entries such that DADA is symmetric. The Lie algebra 𝔤^\hat{\mathfrak{g}} has a Chevalley-Serre presentation given by generators ei,fi,hie_{i},f_{i},h_{i} for 0iN0\leq i\leq N and dd which are subject to the defining relations:

[hi,hj]=0[d,hi]=0[hi,ej]=aijej[hi,fj]=aijfj[h_{i},h_{j}]=0\quad[d,h_{i}]=0\quad[h_{i},e_{j}]=a_{ij}e_{j}\quad[h_{i},f_{j}]=-a_{ij}f_{j}
[ei,fj]=δi,jhi[d,ei]=δ0,iei[d,fi]=δ0,ifi[e_{i},f_{j}]=\delta_{i,j}h_{i}\quad[d,e_{i}]=\delta_{0,i}e_{i}\quad[d,f_{i}]=-\delta_{0,i}f_{i}
(adei)1aij(ej)=0(adfi)1aij(fj)=0(\hbox{$\operatorname{ad}$}e_{i})^{1-a_{ij}}(e_{j})=0\quad(\hbox{$\operatorname{ad}$}f_{i})^{1-a_{ij}}(f_{j})=0

Let 𝔥^\hat{\mathfrak{h}} be the Cartan subalgebra of 𝔤^\hat{\mathfrak{g}} which is the span of {h0,,hN,d}\{h_{0},\ldots,h_{N},d\}.

Recall that affine Lie algebras are classified into two classes: untwisted and twisted, see [11, Ch. 6-8]. In the untwisted case, 𝔤^\hat{\mathfrak{g}} has a natural realization known as loop space realization which is defined by

𝔤^=𝔤[t,t1]cd\hat{\mathfrak{g}}=\mathfrak{g}\otimes\mathbb{C}[t,t^{-1}]\oplus\mathbb{C}c\oplus\mathbb{C}d

where 𝔤\mathfrak{g} is the simple finite dimensional Lie algebra with Cartan matrix (aij)1i,jN(a_{ij})_{1\leq i,j\leq N}, cc is a central element, dd is a degree derivation such that [d,xtn]=nxtn[d,x\otimes t^{n}]=nx\otimes t^{n} for any x𝔤x\in\mathfrak{g} and nn\in\mathbb{Z} and we have [xtn,ytm]=[x,y]tn+m+δn,mn(x|y)c[x\otimes t^{n},y\otimes t^{m}]=[x,y]\otimes t^{n+m}+\delta_{n,-m}n(x|y)c for all x,y𝔤x,y\in\mathfrak{g}, n,mn,m\in\mathbb{Z} where (|)(-|-) is a symmetric invariant bilinear form on 𝔤\mathfrak{g}.

On the other hand, twisted affine Lie algebras are described as fixed points of automorphisms of untwisted algebras. Concretely, let μ~\tilde{\mu} be an automorphism of order r=2r=2 or r=3r=3 of the Coxeter-Dynkin diagram of 𝔤\mathfrak{g} and let μ¯\overline{\mu} be the corresponding diagram automorphism of 𝔤\mathfrak{g}.

Then μ¯\overline{\mu} can be extended to an automorphism μ\mu on 𝔤^=𝔤[t,t1]cd\hat{\mathfrak{g}}=\mathfrak{g}\otimes\mathbb{C}[t,t^{-1}]\oplus\mathbb{C}c\oplus\mathbb{C}d defined as μ(xtm)=(1)m(μ¯(x)tm)\mu(x\otimes t^{m})=(-1)^{m}(\overline{\mu}(x)\otimes t^{m}), for x𝔤x\in\mathfrak{g}, mm\in\mathbb{Z}, μ(c)=c\mu(c)=c, μ(d)=d\mu(d)=d and extended by linearity. The twisted affine Lie algebra (𝔤^)μ(\hat{\mathfrak{g}})^{\mu} is the subalgebra of fixed points of μ\mu.

For example, when r=2r=2,

(𝔤^)μ=(mμ0t2m)(mμ1t2m+1)cd(\hat{\mathfrak{g}})^{\mu}=\left(\sum_{m\in\mathbb{Z}}\mu_{0}\otimes t^{2m}\right)\oplus\left(\sum_{m\in\mathbb{Z}}\mu_{1}\otimes t^{2m+1}\right)\oplus\mathbb{C}c\otimes\mathbb{C}d

where μ0={x𝔤|μ¯(x)=x}\mu_{0}=\{x\in\mathfrak{g}\ |\ \overline{\mu}(x)=x\} and μ1={x𝔤|μ¯(x)=x}\mu_{1}=\{x\in\mathfrak{g}\ |\ \overline{\mu}(x)=-x\} (see [8]).

2.2. Root datum and closed partitions

Let I0={1,,N}I_{0}=\{1,\ldots,N\} and Δ0\Delta_{0} be the root system of 𝔤\mathfrak{g} with θ\theta being the longest positive root. We denote by Q0Q_{0} and P0P_{0} the root and weight lattices of 𝔤\mathfrak{g}. Let I={0,1,,N}I=\{0,1,\ldots,N\}, Δ\Delta the root system of 𝔤^\hat{\mathfrak{g}} with simple roots Π={α0,α1,,αN}\Pi=\{\alpha_{0},\alpha_{1},\ldots,\alpha_{N}\} and let δ=α0+θ\delta=\alpha_{0}+\theta be the indivisible imaginary root. QQ denotes the root lattice, PP the weight lattice, and Qˇ\check{Q}, Pˇ\check{P} denotes the coroot and coweight lattices, respectively. Δre\Delta^{\textrm{re}} and Δim\Delta^{\textrm{im}} denotes the real and the imaginary sets of roots for Δ\Delta.

A subset SS of Δ\Delta is said to be closed if whenever α,βS\alpha,\beta\in S and α+βΔ\alpha+\beta\in\Delta then α+βS\alpha+\beta\in S. We also say that SS is a closed partition if SS is closed, Δ=S(S)\Delta=S\cup(-S) and S(S)=S\cap(-S)=\emptyset. Closed partitions were classified in [3] and [5] (see also [9] and [10]).

For an untwisted affine Lie algebra 𝔤^\hat{\mathfrak{g}}, there are two interesting closed partitions of the root system Δ\Delta, the standard partition and the natural partition, which give rise to two distinct Borel subalgebras that are not conjugate.

The standard partition is defined by

Δst={α+nδ|αΔ0,n>0}Δ0,+{kδ|k>0}\Delta_{\textrm{st}}=\{\alpha+n\delta\ |\ \alpha\in\Delta_{0}\ ,\ n\in\mathbb{Z}_{>0}\}\cup\Delta_{0,+}\cup\{k\delta\ |\ k\in\mathbb{Z}_{>0}\}

and the natural partition by

Δnat={α+nδ|αΔ0,+,n}{kδ|k>0}\Delta_{\textrm{nat}}=\{\alpha+n\delta\ |\ \alpha\in\Delta_{0,+}\ ,\ n\in\mathbb{Z}\}\cup\{k\delta\ |\ k\in\mathbb{Z}_{>0}\}

The respective Borel subalgebras, called standard Borel subalgebra and natural Borel subalgebra, are defined by

𝔟st=(𝔤t[t])𝔫𝔥cd\mathfrak{b}_{\textrm{st}}=\Big{(}\mathfrak{g}\otimes t\mathbb{C}[t]\Big{)}\oplus\mathfrak{n}\oplus\mathfrak{h}\oplus\mathbb{C}c\oplus\mathbb{C}d

and

𝔟nat=(𝔫[t,t1])(𝔥t[t])𝔥cd\mathfrak{b}_{\textrm{nat}}=\Big{(}\mathfrak{n}\otimes\mathbb{C}[t,t^{-1}]\Big{)}\oplus\Big{(}\mathfrak{h}\otimes t\mathbb{C}[t]\Big{)}\oplus\mathfrak{h}\oplus\mathbb{C}c\oplus\mathbb{C}d

where 𝔫=αΔ0,+𝔤α\mathfrak{n}=\bigoplus_{\alpha\in\Delta_{0,+}}\mathfrak{g}_{\alpha}, is the nilpotent Lie subalgebra of the finite Lie algebra 𝔤\mathfrak{g}.

As already mentioned above, a twisted affine algebra is a fixed point set in 𝔤^\hat{\mathfrak{g}} of a non-trivial symmetry of Chevalley generators and, in this case, 𝔟nat\mathfrak{b}_{\textrm{nat}} is the intersection of the fixed point set with the natural Borel subalgebra of 𝔤^\hat{\mathfrak{g}}. For more details see [9].

In this paper, we are going to work with the natural partition of the root system Δnat\Delta_{\textrm{nat}}.

3. Imaginary Verma modules

Let SS be a closed partition of the root system Δ\Delta. Let 𝔤^\hat{\mathfrak{g}} be the untwisted affine Lie algebra which has, with respect to the partition SS, the triangular decomposition 𝔤^=𝔤^S𝔥^𝔤^S\hat{\mathfrak{g}}=\hat{\mathfrak{g}}_{S}\oplus\hat{\mathfrak{h}}\oplus\hat{\mathfrak{g}}_{-S}, where 𝔤^S=αS𝔤^α\hat{\mathfrak{g}}_{S}=\bigoplus_{\alpha\in S}\hat{\mathfrak{g}}_{\alpha} and 𝔥^=𝔥cd\hat{\mathfrak{h}}=\mathfrak{h}\oplus\mathbb{C}c\oplus\mathbb{C}d is an affine Cartan subalgebra. Let U(𝔤^S)U(\hat{\mathfrak{g}}_{S}) and U(𝔤^S)U(\hat{\mathfrak{g}}_{-S}) be, respectively, the universal enveloping algebras of 𝔤^S\hat{\mathfrak{g}}_{S} and 𝔤^S\hat{\mathfrak{g}}_{-S}.

Let λP\lambda\in P. A weight U(𝔤^)U(\hat{\mathfrak{g}})-module V is called an SS-highest weight module with highest weight λ\lambda if there is some non-zero vector vVv\in V such that:

  • uv=0u\cdot v=0 for all u𝔤^Su\in\hat{\mathfrak{g}}_{S}.

  • hv=λ(h)vh\cdot v=\lambda(h)v for all h𝔥^h\in\hat{\mathfrak{h}}.

  • V=U(𝔤^)vU(𝔤^S)vV=U(\hat{\mathfrak{g}})\cdot v\cong U(\hat{\mathfrak{g}}_{-S})\cdot v.

In what follows, let us consider SS to be the natural closed partition of Δ\Delta, i.e., S=ΔnatS=\Delta_{\textrm{nat}} and so 𝔟nat=𝔤^Δnat𝔥^\mathfrak{b}_{\textrm{nat}}=\hat{\mathfrak{g}}_{\Delta_{\textrm{nat}}}\oplus\hat{\mathfrak{h}}. We make \mathbb{C} into a 1-dimensional U(𝔟nat)U(\mathfrak{b}_{\textrm{nat}})-module by picking a generating vector vv and setting (x+h)v=λ(h)v(x+h)\cdot v=\lambda(h)v, for all x𝔤^Δnatx\in\hat{\mathfrak{g}}_{\Delta_{\textrm{nat}}} and h𝔥^h\in\hat{\mathfrak{h}}. The induced module

M(λ)=U(𝔤^)U(𝔟nat)vU(𝔤^Δnat)vM(\lambda)=U(\hat{\mathfrak{g}})\otimes_{U(\mathfrak{b}_{\textrm{nat}})}\mathbb{C}v\cong U(\hat{\mathfrak{g}}_{-\Delta_{\textrm{nat}}})\otimes\mathbb{C}v

is called an imaginary Verma module with Δnat\Delta_{\textrm{nat}}-highest weight λ\lambda. Equivalently, we can define M(λ)M(\lambda) as follows: Let IΔnat(λ)I_{\Delta_{\textrm{nat}}}(\lambda) the ideal of U(𝔤^)U(\hat{\mathfrak{g}}) generated by eik:=eitke_{ik}:=e_{i}\otimes t^{k}, hil:=hitlh_{il}:=h_{i}\otimes t^{l} for iI0i\in I_{0}, kk\in\mathbb{Z}, l>0l\in\mathbb{Z}_{>0}, and by hiλ(hi)1h_{i}-\lambda(h_{i})\cdot 1, dλ(d)1d-\lambda(d)\cdot 1 and cλ(c)1c-\lambda(c)\cdot 1. Then M(λ)=U(𝔤^)/IΔnat(λ)M(\lambda)=U(\hat{\mathfrak{g}})/I_{\Delta_{\textrm{nat}}}(\lambda).

The main properties of this modules, which hold for any affine Lie algebra, were proved in [6] (see also [7] for more properties on this modules), we summarize them in the following.

Proposition 3.1.

Let λP\lambda\in P and let M(λ)M(\lambda) be the imaginary Verma module of Δnat\Delta_{\textrm{nat}}-highest weight λ\lambda. Then M(λ)M(\lambda) has the following properties:

  1. (1)

    The module M(λ)M(\lambda) is a free U(𝔤^Δnat)U(\hat{\mathfrak{g}}_{-\Delta_{\textrm{nat}}})-module of rank 1 generated by the Δnat\Delta_{\textrm{nat}}-highest weight vector 111\otimes 1 of weight λ\lambda.

  2. (2)

    M(λ)M(\lambda) has a unique maximal submodule.

  3. (3)

    Let VV be a U(𝔤^)U(\hat{\mathfrak{g}})-module generated by some Δnat\Delta_{\textrm{nat}}-highest weight vector vv of weight λ\lambda. Then there exists a unique surjective homomorphism ϕ:M(λ)V\phi:M(\lambda)\to V such that 11v1\otimes 1\mapsto v.

  4. (4)

    dimM(λ)λ=1\dim M(\lambda)_{\lambda}=1. For any μ=λkδ\mu=\lambda-k\delta, k>0k\in\mathbb{Z}_{>0}, 0<dimM(λ)μ<0<\dim M(\lambda)_{\mu}<\infty. If μλkδ\mu\neq\lambda-k\delta for any integer k0k\geq 0 and M(λ)μ0M(\lambda)_{\mu}\neq 0, then dimM(λ)μ=\dim M(\lambda)_{\mu}=\infty.

  5. (5)

    Let λ,μ𝔥^\lambda,\mu\in\hat{\mathfrak{h}}^{*}. Any non-zero element of HomU(𝔤^)(M(λ),M(μ))\operatorname{Hom}_{U(\hat{\mathfrak{g}})}(M(\lambda),M(\mu)) is injective.

  6. (6)

    The module M(λ)M(\lambda) is irreducible if and only if λ(c)0\lambda(c)\neq 0.

Suppose now that λ(c)=0\lambda(c)=0 and consider the ideal JΔnat(λ)J_{\Delta_{\textrm{nat}}}(\lambda) generated by IΔnat(λ)I_{\Delta_{\textrm{nat}}}(\lambda) and hilh_{il}, iI0i\in I_{0} and l{0}l\in\mathbb{Z}\setminus\{0\}. Set

M~(λ)=U(𝔤^)/JΔnat(λ)\tilde{M}(\lambda)=U(\hat{\mathfrak{g}})/J_{\Delta_{\textrm{nat}}}(\lambda)

Then M~(λ)\tilde{M}(\lambda) is a homomorphic image of M(λ)M(\lambda) which we call reduced imaginary Verma module. The following is proved in [6], Theorem 1.

Proposition 3.2.

M~(λ)\tilde{M}(\lambda) is irreducible if and only if λ(hi)0\lambda(h_{i})\neq 0 for all iI0i\in I_{0}.

4. The category 𝒪red,im\mathcal{O}_{red,im}

Consider the Heisenberg subalgebra GG which by definition is

G=k{0}𝔤^kδcG=\bigoplus_{k\in\mathbb{Z}\setminus\{0\}}\hat{\mathfrak{g}}_{k\delta}\oplus\mathbb{C}c

We will say that a 𝔤^\hat{\mathfrak{g}}-module V is GG-compatible if:

  1. (i)

    VV has a decomposition V=T(V)TF(V)V=T(V)\oplus TF(V) where T(V)T(V) and TF(V)TF(V) are non-zero GG-modules, called, respectively, torsion and torsion free module associated to VV.

  2. (ii)

    himh_{im} for iI0i\in I_{0}, m{0}m\in\mathbb{Z}\setminus\{0\} acts bijectively on TF(V)TF(V), i.e., they are bijections on TF(V)TF(V).

  3. (iii)

    TF(V)TF(V) has no non-zero 𝔤^\hat{\mathfrak{g}}-submodules.

  4. (iv)

    GT(V)=0G\cdot T(V)=0.

Consider the set

𝔥^red={λ𝔥^|λ(c)=0,λ(hi)0 for any iI0}\hat{\mathfrak{h}}^{*}_{red}=\{\lambda\in\hat{\mathfrak{h}}^{*}\;|\;\lambda(c)=0,\lambda(h_{i})\notin\mathbb{Z}_{\geq 0}\mbox{ for any }i\in I_{0}\}

We define the category 𝒪red,im\mathcal{O}_{red,im} as the category whose objects are 𝔤^\hat{\mathfrak{g}}-modules MM such that

  1. (1)

    MM is 𝔥^red\hat{\mathfrak{h}}^{*}_{red}-diagonalizable, that means,

    M=ν𝔥^redMν, where Mν={mM|him=ν(hi)m,dm=ν(d)m,iI0}M=\bigoplus_{\nu\in\hat{\mathfrak{h}}^{*}_{red}}M_{\nu},\mbox{ where }M_{\nu}=\{m\in M|h_{i}m=\nu(h_{i})m,dm=\nu(d)m,i\in I_{0}\}
  2. (2)

    For any iI0i\in I_{0} and any nn\in\mathbb{Z}, eine_{in} acts locally nilpotently.

  3. (3)

    MM is GG-compatible.

  4. (4)

    The morphisms between modules are 𝔤^\hat{\mathfrak{g}}-homomorphisms

Example 4.1.

Reduced imaginary Verma modules belongs to 𝒪red,im\mathcal{O}_{red,im}. Indeed, for M~(λ)\tilde{M}(\lambda) consider T(M~(λ))=vλT(\tilde{M}(\lambda))=\mathbb{C}v_{\lambda} and TF(V)=k,n1,,nN0M~(λ)λ+kδn1α1nNαNTF(V)=\bigoplus_{k\in\mathbb{Z},n_{1},\ldots,n_{N}\in\mathbb{Z}_{\geq 0}}\tilde{M}(\lambda)_{\lambda+k\delta-n_{1}\alpha_{1}-\ldots-n_{N}\alpha_{N}}, and at least one nj0n_{j}\neq 0. Moreover, direct sums of reduced imaginary Verma modules belongs to 𝒪red,im\mathcal{O}_{red,im}.

Recall that a loop module for 𝔤^\hat{\mathfrak{g}} is any representation of the form M^:=M[t,t1]\hat{M}:=M\otimes\mathbb{C}[t,t^{-1}] where MM is a 𝔤\mathfrak{g}-module and the action of 𝔤^\hat{\mathfrak{g}} on M^\hat{M} is given by

(xtk)(mtl):=(xm)tk+l,c(mtl)=0(x\otimes t^{k})(m\otimes t^{l}):=(x\cdot m)\otimes t^{k+l}\quad,\quad c(m\otimes t^{l})=0

for x𝔤x\in\mathfrak{g}, mMm\in M and k,lk,l\in\mathbb{Z}. Here xmx\cdot m is the action of x𝔤x\in\mathfrak{g} on mMm\in M.

Proposition 4.2.

Let MM is a 𝔤\mathfrak{g}-module in the BGG category 𝒪\mathcal{O}. Then the loop module M^\hat{M} can not lie in 𝒪red,im\mathcal{O}_{red,im}.

Proof. Let M𝒪M\in\mathcal{O} and let M^\hat{M} be its associated loop module. If MM is finite dimensional, it is a direct sum of finite dimensional irreducible 𝔤\mathfrak{g}-modules, and these have highest weights which are non-negative integers when evaluated in hih_{i} for any iI0i\in I_{0}. So, condition (1) is not satisfied and M^\hat{M} does not belongs to 𝒪red,im\mathcal{O}_{red,im}. Assume now that MM is an infinite dimensional 𝔤\mathfrak{g}-module. Note that condition (2) is satisfied as 𝔫\mathfrak{n} acts locally nilpotently on MM. If condition (1) does not hold, we are done. Suppose that (1) holds and that M^\hat{M} is GG-compatible. We have M^=T(M^)TF(M^)\hat{M}=T(\hat{M})\oplus TF(\hat{M}) satisfying (i) - (iv) above. Take any nonzero element i=kkmitiT(M^)\sum_{i=-k}^{k}m_{i}\otimes t^{i}\in T(\hat{M}) with miMμm_{i}\in M_{\mu} for some weight μ¯𝔥^red\bar{\mu}\in\hat{\mathfrak{h}}^{*}_{red}. Then by (iv) we have

0=(hjtr)(i=kkmiti)=i=kk(hjmi)ti+r=μ¯(hj)(i=kkmiti+r)0=(h_{j}\otimes t^{r})\left(\sum_{i=-k}^{k}m_{i}\otimes t^{i}\right)=\sum_{i=-k}^{k}(h_{j}\cdot m_{i})\otimes t^{i+r}=\bar{\mu}(h_{j})\left(\sum_{i=-k}^{k}m_{i}\otimes t^{i+r}\right)

where jI0j\in I_{0}, r{0}r\in\mathbb{Z}\setminus\{0\}. Hence μ¯(hj)=0\bar{\mu}(h_{j})=0, for any jI0j\in I_{0}, which contradicts to the fact that μ¯𝔥^red\bar{\mu}\in\hat{\mathfrak{h}}^{*}_{red}. Then T(M^)=0T(\hat{M})=0 and M^=TF(M^)\hat{M}=TF(\hat{M}) which is a 𝔤^\hat{\mathfrak{g}}-module contradicting (i) and (iii), and thus (3). This completes the proof.

\Box

5. Main results

In this section we will show that the category 𝒪red,im\mathcal{O}_{red,im} is a semisimple category having reduced imaginary Verma modules as its simple objects. First we will show that reduced imaginary Verma modules have no nontrivial extensions in 𝒪red,im\mathcal{O}_{red,im}.

Theorem 5.1.

If λ,μ𝔥^red\lambda,\mu\in\hat{\mathfrak{h}}^{*}_{red} then Ext𝒪red,im1(M~(λ),M~(μ))=0\operatorname{Ext}_{\mathcal{O}_{red,im}}^{1}(\tilde{M}(\lambda),\tilde{M}(\mu))=0.

Proof. Let MM be an extension of M~(λ)\tilde{M}(\lambda) and M~(μ)\tilde{M}(\mu) that fits in the following short exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M~(λ)\textstyle{\tilde{M}(\lambda)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}M~(μ)\textstyle{\tilde{M}(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Suppose μ=λ+kδi=1Nsiαi\mu=\lambda+k\delta-\sum_{i=1}^{N}s_{i}\alpha_{i}, for sis_{i}\in\mathbb{Z} and kk\in\mathbb{Z}, and all sis_{i}’s have the same sign or equal to 0. First, consider the case when si=0s_{i}=0 for all iI0i\in I_{0}. Then μ=λ+kδ\mu=\lambda+k\delta and so, in MM there will be two vectors vλv_{\lambda} and vμv_{\mu} of weights λ\lambda and μ\mu respectively, annihilated by 𝔫[t,t1]\mathfrak{n}\otimes\mathbb{C}[t,t^{-1}]. Moreover, because of the condition (iv) in the definition of GG-compatibility, these two points are isolated. So, vλv_{\lambda} and vμv_{\mu} are highest weight vectors, each of which generates an irreducible subrepresentation (isomorphic to M~(λ)\tilde{M}(\lambda) and M~(μ)\tilde{M}(\mu) respectively), and the extension splits. Hence, we can assume that not all sis_{i} are equal to zero and that the map ι:M~(λ)M\iota:\tilde{M}(\lambda)\to M in the short exact sequence is an inclusion. Assume that si0s_{i}\in\mathbb{Z}_{\geq 0} for all ii.

Let v¯μM\overline{v}_{\mu}\in M be a preimage under the map π\pi of a highest weight vector vμM~(μ)v_{\mu}\in\tilde{M}(\mu) of weight μ\mu. We have (𝔫[t,t1])vμ=Gvμ=0(\mathfrak{n}\otimes\mathbb{C}[t,t^{-1}])v_{\mu}=Gv_{\mu}=0, and we are going to show that Gv¯μ=0G\overline{v}_{\mu}=0. Assume that v¯μT(M)\overline{v}_{\mu}\notin T(M). Then we claim that T(M)=vλT(M)=\mathbb{C}v_{\lambda}. Indeed, we have vλT(M)\mathbb{C}v_{\lambda}\subset T(M). If uT(M)vλu\in T(M)\setminus\mathbb{C}v_{\lambda} is some nonzero weight element, then Gu=0G\cdot u=0 and π(u)\pi(u) belongs to T(M~(μ))=vμT(\tilde{M}(\mu))=\mathbb{C}v_{\mu}. If π(u)=0\pi(u)=0 then uM~(λ)u\in\tilde{M}(\lambda) which is a contradiction. If π(u)\pi(u) is a nonzero multiple of vμv_{\mu}, then uu has weight μ\mu and thus uu is a multiple of v¯μ\overline{v}_{\mu} which is again a contradiction. So, we assume T(M)=vλT(M)=\mathbb{C}v_{\lambda}.

Note that for any iI0i\in I_{0} and m{0}m\in\mathbb{Z}\setminus\{0\} we have

π(himv¯μ)=himπ(v¯μ)=himvμ=0.\pi(h_{im}\overline{v}_{\mu})=h_{im}\pi(\overline{v}_{\mu})=h_{im}v_{\mu}=0.

Then himv¯μM~(λ)h_{im}\overline{v}_{\mu}\in\tilde{M}(\lambda). Suppose there exists jI0j\in I_{0} such that hjmv¯μ0h_{jm}\overline{v}_{\mu}\neq 0 for m{0}m\in\mathbb{Z}\setminus\{0\}. Because hjmv¯μM~(λ)h_{jm}\overline{v}_{\mu}\in\tilde{M}(\lambda) and has weight μ+mδ\mu+m\delta, it belongs to TF(M~(λ))TF(\tilde{M}(\lambda)). Hence, there exists a nonzero vM~(λ)v^{\prime}\in\tilde{M}(\lambda) of weight μ\mu such that hjmv¯μ=hjmvh_{jm}\overline{v}_{\mu}=h_{jm}v^{\prime}. Hence, hjm(v¯μv)=0h_{jm}(\overline{v}_{\mu}-v^{\prime})=0 implying v¯μvT(M)vλ\overline{v}_{\mu}-v^{\prime}\in T(M)\cong\mathbb{C}v_{\lambda}. Then v¯μv=pvλ\overline{v}_{\mu}-v^{\prime}=p\ v_{\lambda}, for some pp\in\mathbb{C}. Comparing the weight we arrive to a contradiction. Hence, hinv¯μ=0h_{in}\overline{v}_{\mu}=0. So, we get Gv¯μ=0G\overline{v}_{\mu}=0.

Recall that the operators eime_{im} acts locally nilpotently on M~(λ)\tilde{M}(\lambda). We claim that eimv¯μ=0e_{im}\overline{v}_{\mu}=0 for all possible ii and nn. Indeed, assume that ejmv¯μ0e_{jm}\overline{v}_{\mu}\neq 0 for some jI0j\in I_{0} and some integer mm. Then eimv¯μM~(λ)e_{im}\overline{v}_{\mu}\in\tilde{M}(\lambda). Consider the 𝔰𝔩^2\hat{\mathfrak{sl}}_{2}-subalgebra 𝔰(j)\mathfrak{s}(j) generated by fjn,ejnf_{jn},e_{jn} and hjlh_{jl} for n,ln,l\in\mathbb{Z}. Let MjM_{j} be an 𝔰(j)\mathfrak{s}(j)-submodule of MM generated by v¯μ\overline{v}_{\mu}. Then MjM_{j} is an extension of reduced imaginary Verma 𝔰(j)\mathfrak{s}(j)-modules, one of which of highest weight μ\mu. Since M𝒪red,imM\in\mathcal{O}_{red,im}, we immediately see that MjM_{j} is an object of the corresponding reduced category 𝒪red,im(𝔰(j))\mathcal{O}_{red,im}(\mathfrak{s}(j)) for 𝔰(j)\mathfrak{s}(j). But this category is semisimple by [2]. Hence, eimv¯μ=0e_{im}\overline{v}_{\mu}=0 for all ii and mm. Therefore, v¯μ\overline{v}_{\mu} generates a 𝔤\mathfrak{g}-submodule of MM isomorphic to M~(μ)\tilde{M}(\mu) and the short exact sequence splits.

Assume now that si0s_{i}\in\mathbb{Z}_{\leq 0} for all ii and not all of them are 0. As M~(μ)\tilde{M}(\mu) is irreducible and M~(λ)\tilde{M}(\lambda) is a 𝔤\mathfrak{g}-submodule of MM, the short exact sequence splits completing the proof.

\Box

Remark 5.2.

Observe that modules M~(λ)\tilde{M}(\lambda) and M~(λkδ)\tilde{M}(\lambda-k\delta) have a nontrivial extension in the category of 𝔤^\hat{\mathfrak{g}}-modules for any integer kk.

Theorem 5.3.

If MM is an irreducible module in the category 𝒪red,im\mathcal{O}_{red,im}, then MM~(λ)M\cong\tilde{M}(\lambda) for some λ𝔥^red\lambda\in\hat{\mathfrak{h}}_{red}^{*}.

Proof. Let MM be an irreducible module in 𝒪red,im\mathcal{O}_{red,im}. As a GG-module, MT(M)TF(M)M\cong T(M)\oplus TF(M) where both summands are non-zero. Let vT(M)v\in T(M) be a non-zero element of weigh λ𝔥^red\lambda\in\hat{\mathfrak{h}}_{red}^{*}. Then himv=0h_{im}v=0 for all iI0i\in I_{0} and all m{0}m\in\mathbb{Z}\setminus\{0\}. For each iI0i\in I_{0} let pi>0p_{i}\in\mathbb{Z}_{>0} be the minimum possible integer such that ei0piv=0e_{i0}^{p_{i}}v=0. If all pi=1p_{i}=1 we have ei0v=0e_{i0}v=0 and then, because [hin,ei0]=2ein[h_{in},e_{i0}]=2e_{in} we get that einv=0e_{in}v=0 for all iI0i\in I_{0} and n{0}n\in\mathbb{Z}\setminus\{0\}. Hence, we have an epimorphism M~(λ)M\tilde{M}(\lambda)\twoheadrightarrow M, since λ𝔥^red\lambda\in\hat{\mathfrak{h}}_{red}^{*}, M~(λ)\tilde{M}(\lambda) is simple and so MM~(λ)M\cong\tilde{M}(\lambda).

On the other hand, assume there exists at least one pip_{i} such that pi>1p_{i}>1. We are going to construct a set of elements in MM which are killed by ei0e_{i0} for all iI0i\in I_{0}. First of all, set p(1)=max{pi|iI0}p^{(1)}=\max\{p_{i}|i\in I_{0}\} and set wi:=ei0p(1)1vw_{i}:=e_{i0}^{p^{(1)}-1}v. Note that wi=0w_{i}=0 if p(1)>pip^{(1)}>p_{i} and wi0w_{i}\neq 0 if p(1)=pip^{(1)}=p_{i}, so at least one wiw_{i} in non-zero. If for all jI0j\in I_{0}, ej0wi=0e_{j0}w_{i}=0 we are done, if not there exists numbers pij>0p_{ij}\in\mathbb{Z}_{>0} such that ej0pijwi=0e_{j0}^{p_{ij}}w_{i}=0 and some of the pijp_{ij} are strictly bigger than 11. Set p(2)=max{pij|i,jI0}p^{(2)}=\max\{p_{ij}|i,j\in I_{0}\} and set wij=ej0p(2)1wiw_{ij}=e_{j0}^{p^{(2)}-1}w_{i}, note that at least one wijw_{ij} is non-zero. If ek0wij=0e_{k0}w_{ij}=0 for all kI0k\in I_{0} we are done, if not we repeat the process. Because of the locally nilpotency of the el0e_{l0} for lI0l\in I_{0}, in finitely many steps, let say \ell steps, we can find at least one non-zero element w𝐢w_{\bf i}, for 𝐢=i1i2i{\bf i}=i_{1}i_{2}\ldots i_{\ell} a string of elements in I0I_{0} such that el0w𝐢=0e_{l0}w_{\bf i}=0. Moreover, if 𝐢{\bf i}^{-} denotes the string i1i2i1i_{1}i_{2}\ldots i_{\ell-1}, then w𝐢=ei0p()1w𝐢w_{\bf i}=e_{i_{\ell}0}^{p^{(\ell)}-1}w_{{\bf i}^{-}} and so, for all n{0}n\in\mathbb{Z}\setminus\{0\}, 0=hinei0p()w𝐢=2p()einw𝐢0=h_{i_{\ell}n}e_{i_{\ell}0}^{p^{(\ell)}}w_{{\bf i}^{-}}=2p^{(\ell)}e_{i_{\ell}n}w_{\bf i}, i.e., einw𝐢=0e_{i_{\ell}n}w_{\bf i}=0. Now, 0=hj0ejmei0p()w𝐢=ejmhj0ei0p()w𝐢+2ejmei0p()w𝐢=2p()ejmei0p()1w𝐢=2p()ejmw𝐢0=h_{j0}e_{jm}e_{i_{\ell}0}^{p^{(\ell)}}w_{{\bf i}^{-}}=e_{jm}h_{j0}e_{i_{\ell}0}^{p^{(\ell)}}w_{{\bf i}^{-}}+2e_{jm}e_{i_{\ell}0}^{p^{(\ell)}}w_{{\bf i}^{-}}=2p^{(\ell)}e_{jm}e_{i_{\ell}0}^{p^{(\ell)}-1}w_{{\bf i}^{-}}=2p^{(\ell)}e_{jm}w_{\bf i}.

Pick one of the non-zero w𝐢w_{\bf i} constructed above and let W𝐢=U(G)w𝐢W_{\bf i}=U(G)w_{\bf i} be a GG-submodule of MM. By construction elnW𝐢=0e_{ln}W_{\bf i}=0 for all lI0l\in I_{0} and nn\in\mathbb{Z}. Considered the induced module I(W𝐢)=IndGHN+𝔤^W𝐢I(W_{\bf i})=\operatorname{Ind}_{G\oplus H\oplus N_{+}}^{\hat{\mathfrak{g}}}W_{\bf i}, where N+=iI0,nZeinN_{+}=\bigoplus_{i\in I_{0},n\in Z}\mathbb{C}e_{in} acts by 0, H=iI0hidH=\bigoplus_{i\in I_{0}}\mathbb{C}h_{i}\oplus\mathbb{C}d acts by hiw𝐢=μ(hi)w𝐢h_{i}w_{\bf i}=\mu(h_{i})w_{\bf i}, dw𝐢=μ(d)w𝐢dw_{\bf i}=\mu(d)w_{\bf i}, for some weight μ\mu. Because MM is simple, it is a quotient of I(W𝐢)I(W_{\bf i}). If w𝐢T(M)w_{\bf i}\in T(M), we have W𝐢=w𝐢W_{\bf i}=\mathbb{C}w_{\bf i}, and so MM is a quotient of I(W𝐢)=M~(λ)I(W_{\bf i})=\tilde{M}(\lambda) and we are done.

In case w𝐢T(M)w_{\bf i}\notin T(M), as in the proof of Proposition 6.0.3. of [2] we get a contradiction. This completes the proof.

\Box

Proposition 5.4.

If MM is an arbitrary object in 𝒪red,im\mathcal{O}_{red,im}, then Mλi𝔥^redM~(λi)M\cong\bigoplus_{\lambda_{i}\in\hat{\mathfrak{h}}^{*}_{red}}\tilde{M}(\lambda_{i}), for some λis\lambda_{i}^{\prime}s.

Proof. Because MM is in 𝒪red,im\mathcal{O}_{red,im}, it is a GG-compatible and so, it has a decomposition as a GG-module given by MT(M)TF(M)M\cong T(M)\oplus TF(M). Since all the weights of MM are in 𝔥^red\hat{\mathfrak{h}}^{*}_{red}, T(M)T(M) is not a 𝔤^\hat{\mathfrak{g}}-submodule of MM. Indeed, suppose T(M)T(M) is a 𝔤^\hat{\mathfrak{g}}-module. let vT(M)v\in T(M) and consider f0vT(M)f_{0}v\in T(M). Then h0mf0v=0h_{0m}f_{0}v=0 and fmv=0f_{m}v=0 for any m0m\neq 0. Applying h0,mh_{0,-m} we get h0,mfmv=0h_{0,-m}f_{m}v=0 and f0v=0f_{0}v=0. Since the weight of vv is in 𝔥^red\hat{\mathfrak{h}}^{*}_{red}, e0pv0e_{0}^{p}v\neq 0 for any p>0p>0. But if pp is sufficiently large the weigh of e0pve_{0}^{p}v will not be in 𝔥^red\hat{\mathfrak{h}}^{*}_{red} and we get a contradiction.

Let vT(M)v\in T(M) non-zero. As in the proof of the previous statement there exists a string 𝐢{\bf i} of elements of I0I_{0} and a vector w𝐢w_{\bf i} such that ejmw𝐢=0e_{jm}w_{\bf i}=0 for all jI0j\in I_{0} and mm\in\mathbb{Z}. Let W𝐢=U(G)w𝐢W_{\bf i}=U(G)w_{\bf i}. Then we have two possibilities: either w𝐢T(M)w_{\bf i}\notin T(M) or w𝐢T(M)w_{\bf i}\in T(M).

In the first case, consider the induced module I(W𝐢)I(W_{\bf i}). Clearly TF(I(W𝐢))I(W𝐢)TF(I(W_{\bf i}))\subseteq I(W_{\bf i}). Now, if wI(W𝐢)w\in I(W_{\bf i}), because w𝐢T(M)w_{\bf i}\notin T(M) we have gw0gw\neq 0 for gGg\in G and so wTF(I(W𝐢))w\in TF(I(W_{\bf i})). Then TF(I(W𝐢))=I(W𝐢)TF(I(W_{\bf i}))=I(W_{\bf i}). By the five lemma, any quotient and subquotient of I(W𝐢)I(W_{\bf i}) also satisfies this property. Set M:=U(𝔤^)w𝐢M^{\prime}:=U(\hat{\mathfrak{g}})w_{\bf i} which is a subquotient of I(W𝐢)I(W_{\bf i}). Then MM^{\prime} is a 𝔤^\hat{\mathfrak{g}}-submodule of MM and so M=TF(M)M^{\prime}=TF(M^{\prime}) is a 𝔤^\hat{\mathfrak{g}}-submodule of TF(M)TF(M), but TF(M)TF(M) does not have proper 𝔤^\hat{\mathfrak{g}}-submodule and so M=TF(M)M^{\prime}=TF(M). But, W𝐢W_{\bf i} is a proper GG-submodule of MM^{\prime} which is not possible because MM is in 𝒪red,im\mathcal{O}_{red,im}. And so, this case does not occur.

In the second case, W𝐢=w𝐢T(M)W_{\bf i}=\mathbb{C}w_{\bf i}\subseteq T(M). So, as 𝔤^\hat{\mathfrak{g}}-modules I(W𝐢)M~(λ𝐢)I(W_{\bf i})\cong\tilde{M}(\lambda_{\bf i}) for some λ𝐢\lambda_{\bf i}, is a 𝔤^\hat{\mathfrak{g}}-submodule of MM. Then, any non-zero element of T(M)T(M) generates an irreducible reduced imaginary Verma module which is a 𝔤^\hat{\mathfrak{g}}-submodule of MM and because there are no extensions between them, they are direct summands on MM.

\Box

Corollary 5.5.

The category 𝒪red,im\mathcal{O}_{red,im} is closed under taking subquotients and direct sums, so it is a Serre subcategory.

Remark 5.6.

The proofs on the above statements depends on the structure of reduced imaginary Verma modules, the closed partition Δnat\Delta_{\textrm{nat}} and the associated Borel subalgebra 𝔟nat{\mathfrak{b}}_{\textrm{nat}}. But, the properties of reduced imaginary Verma modules hold for both untwisted or twisted affine Lie algebras. Moreover, the natural Borel subalgebra for the twisted Lie algebra is properly contained in the natural Borel subalgebra for the untwisted case. So, the results above hold for any affine Lie algebra.

Acknowledgement

JCA has been support by the FAPESP Grant 2021/13022-9.

References

  • [1] B. Cox. Verma modules induced from nonstandard Borel subalgebras. Pacific J. Math., 165(2):269–294, 1994.
  • [2] B. Cox, V. Futorny, and K. Misra. Imaginary Verma modules for Uq(𝔰𝔩(2)^)U_{q}(\widehat{\mathfrak{sl}(2)}) and crystal-like bases. J. Algebra, 481:12–35, 2017.
  • [3] V. Futorny. Parabolic partitions of root systems and corresponding representations of the affine Lie algebras. Akad. Nauk Ukrain. SSR Inst. Mat. Preprint, (8):30–39, 1990.
  • [4] V. Futorny. On imaginary verma modules over the affine lie algebra A1(1){A}_{1}^{(1)}. Preprint series: Pure mathematics http://urn. nb. no/URN: NBN: no-8076, 1991.
  • [5] V. Futorny. The parabolic subsets of root system and corresponding representations of affine Lie algebras. In Proceedings of the International Conference on Algebra, Part 2 (Novosibirsk, 1989), volume 131 of Contemp. Math., pages 45–52. Amer. Math. Soc., Providence, RI, 1992.
  • [6] V. Futorny. Imaginary Verma modules for affine Lie algebras. Canad. Math. Bull., 37(2):213–218, 1994.
  • [7] V. Futorny. Irreducible non-dense A1(1){A}_{1}^{(1)}-modules. Pacific Journal of Mathematics, 172(1):83–99, 1996.
  • [8] V. Futorny. Representations of affine Lie algebras. Queen’s Papers in Pure and Applied Math., 106, 1997.
  • [9] H. Jakobsen and V. Kac. A new class of unitarizable highest weight representations of infinite-dimensional Lie algebras. In Nonlinear equations in classical and quantum field theory (Meudon/Paris, 1983/1984), volume 226 of Lecture Notes in Phys., pages 1–20. Springer, Berlin, 1985.
  • [10] H. Jakobsen and V. Kac. A new class of unitarizable highest weight representations of infinite-dimensional Lie algebras. II. J. Funct. Anal., 82(1):69–90, 1989.
  • [11] V. Kac. Infinite-dimensional Lie algebras. Cambridge University Press, Cambridge, third edition, 1990.